Sie sind auf Seite 1von 0

Hans Ulrich Sss

Norbert Franz Nimmerfroh


Degussa AG
HYDROGEN PEROXI DE I N HYDROGEN PEROXI DE I N
CHEMI CAL PULP BLEACHI NG CHEMI CAL PULP BLEACHI NG
-an overview
Paper for the ABTCP meeting on pulp bleaching at
Vitoria, Espirito Santo, Brasil,
June 25. - 28, 1996
1. Abstract
If applied under appropriate conditions, hydrogen
peroxide is a very effective bleaching chemical. The
very wide application range will be described, starting
with boosting the brightness in an existing bleach
plant in any extraction stage, as well as in an
additional final treatment in the high density storage
tower. In addition hydrogen peroxide decreases the
demand for chlorine dioxide in the final bleaching
stages of an ECF sequence. Significant savings are
the result of peroxide addition in the second extrac-
tion stage. The replacement of one chlorine dioxide
stage in D
1
-D
2
final bleaching with peroxide and the
application of a D
1
-P treatment also cuts the active
chlorine demand.
With hydrogen peroxide the kappa number in the first
extraction stage can be lowered and the substitution
of chlorine with chlorine dioxide made more efficient.
In TCF sequences hydrogen peroxide is the chemical
of choice to gain the required brightness. The
conditions of peroxide bleaching are described and
the benefit of magnesium salt addition and buffering
with very small amounts of sodium silicate is dem-
onstrated.
In addition the effect of alkaline hydrogen peroxide on
the corrosion of titanium is reviewed and underlined
with laboratory data. Basically, corrosion does not
take place under the typical reaction conditions for
peroxide bleaching in an ECF sequence.
2. Introduction
In 1995 the worldwide production of hydrogen per-
oxide was estimated to have reached more than 1,6
million metric tons. The dominating amount of hy-
drogen peroxide is used in bleaching processes.
Hydrogen peroxide is generally supplied as an
aqueous solution, typically in concentrations between
35% by weight and up to 70% by weight. These
acidic solutions of hydrogen peroxide in water are
very stable. Hydrogen peroxide can be stored for
months in stainless steel tanks without significant
changes of the content.
Compared to water, the energy content of hydrogen
peroxide is much higher. For water the heat of for-
mation (H, eq. 1) from the elements is as low as -
286 kJ mol
-1
, for hydrogen peroxide (eq. 2) the co-
responding value is only -188 kJ mol
-1
[1]. In conse-
quence hydrogen peroxide is less stable and can
disproportionate into water and oxygen.
(1) H
2
+ O
2
H
2
O H = -286 kJ mol
-1
(2) H
2
+ O
2
H
2
O
2
H = -188 kJ mol
-1
Since the activation energy for the cleavage of the
oxygen-oxygen bond is rather low (H = -71 kJ mol
-1
)
[2], traces of contaminants can start this reaction.
Basically the decomposition is a redox process.
Hydrogen peroxide either supplies electrons and
yields oxygen or accepts electrons and yields water.
Metal salts of each state of oxidation can start the
decomposition reaction. The first step can be the
reduction of iron according to equation (3):
(3) 2 Me
2+
+ H
2
O
2
2 Me
+
+ O
2
+ 2 H
+
The alternative is the oxidation of a metal according
to equation (4):
(4) 2 Me
+
+ H
2
O
2
+ 2 H
+
2 Me
2+
+ 2 H
2
O
The reaction certainly can also start with the reduced
form of metal. The overall reaction is identical, it is
the formation of water and oxygen from hydrogen
peroxide and the redox system of the metal is acting
as the catalyst [1].
The decomposition of hydrogen peroxide is in addition
catalysed by alkali. These are the reaction steps:
(5) H
2
O
2
+ OH
-
H
2
O + HOO
-
(6) HOO
-
+ H
2
O
2
H
2
O + O
2
+ OH
-
Since bleaching with hydrogen peroxide requires
alkaline conditions, this decomposition reaction is
very important for the practice.
Single electron transfer reactions of hydrogen per-
oxide with catalysts yield radicals. These decom-
position reactions take place with metals or enzymes
like catalase. Radicals formation can also be the
result of a thermal cleavage of the oxygen-oxygen
bond.
(7) H
2
O
2
+ Me
+
OH
-
+
.
OH + Me
2+
(8) H
2
O
2
+
.
OH H
2
O +
.
OOH
(9) H
2
O +
.
OOH
.
OO
-
+ H
3
O
+
2
The hydroxyl radical, the hydroperoxy radical and the
superoxide anion radical are important intermediates.
They cause side reactions in bleaching processes,
with delignification as positive and depolymerisation
of the cellulose as negative result. Generally radicals
produce more negative effects than positive results on
delignification. Therefore, transition metal ions are
normally eliminated before a peroxide treatment.
In pulp bleaching the formation of the perhydroxyl
anion (eq. 5), a nucleophile intermediate, is re-
sponsible for the oxidation of chromophores in lignin
through the cleavage of side chains [3]. The effect of
a hydrogen peroxide treatment is dominantly an
increase of the brightness. Delignification with
hydrogen peroxide is to a large extent the result of
the action of the radicals produced in equations 7, 8
and 9 [3, 4]. At moderate temperature, under buffered
conditions and in the absence of transition metals,
the delignifying effect of hydrogen peroxide is limited.
The degradation of polymerized lignin, which can be
the result of high intensity pulping conditions, with
hydrogen peroxide is limited. The perhydroxyl anion,
being a nucleophile, cannot attack the electron rich
aromatic rings of the residual lignin.
3. Application of hydrogen peroxide on mill
scale
Bleaching reactions with hydrogen peroxide are
conducted in stainless steel reactors or in rubber
coated carbon steel towers, to avoid decomposition of
the peroxide and corrosion of the reactors. Alkaline
peroxide solutions are not corroding stainless steel.
Peroxide solutions are diluted and premixed with
caustic soda before they are added to the pulp. The
alternative is the addition of the technical grade
solution of hydrogen peroxide (50% to 70%) to the
pulp without dilution, which requires an efficient mixer
in order to avoid local overconcentration. The
advantage is the addition of a lower volume of diluting
liquid which permits a higher consistency. Generally
the consistency in a peroxide bleaching stage should
be as high as possible. The increase of the relative
concentration of the chemicals to the pulp and the
elimination of dissolved organic compounds improves
the efficency of the bleaching.
Very alkaline peroxide solutions are agressive against
another material which is widely used to in the
construction of bleach plants: titanium. The corrosion
of titanium through alkaline peroxide solutions is
strongly depending on the peroxide concentration and
the pH value. Table 1 and figure 1 have the results of
laboratory tests with titanium metal treated with
alkaline hydrogen peroxide at 80C.
Table 1:
Corrosion of titanium with alkaline hydrogen peroxide.
Effect of peroxide concentration and pH value. A
continuous flow of peroxide and alkali was applied to
maintain the concentration of the chemicals, the
amount of chemicals is calculated on liquid, it has to
be multiplied by ten to correspond (at 10%
consistency) to pulp.
H
2
O
2
(%)
NaOH
(%)
Na
2
CO
3
(%)
resultin
g pH
corrosion
rate
(mm/year)
0,2 - 0,1 10,3 0
0,2 - 0,2 10,6 0
0,2 0,1 0,1 10,8 0
0,2 0,2 - 11,6 0,06
0,2 0,4 - 12,0 0,22
0,5 0,2 - 11,6 0,28
0,5 0,4 - 11,9 1,62
Na2CO3 %: 0,1 0,2 0,1 - -
0,2 0,2 0,2 0,2 0,2
10
10,5
11
11,5
12
12,5
pH value
0
0,5
1
1,5
2
2,5
corrosion (mm/year)
resulting pH
corrosion rate
H2O2 %:
NaOH %: - - 0,1 0,2 0,4
Figure 1:
Corrosion of titanium with alkaline peroxide, effect of
pH level
The results indicate a significant increase of the
corrosion level at high pH-values. In addition, the
losses of material increase with increasing hydrogen
peroxide concentration. On the other hand, at a pH
below pH 11, no corrosion takes place. Hydrogen
peroxide starts the corrosion only at higher concen-
tration levels. The highest corrosion rate in the table
coresponds to a hydrogen peroxide charge of 5 %
H
2
O
2
on pulp at 10% consistency, an addition value,
which is far away from typical peroxide charges. The
same is valid for the charges of caustic soda. These
are typically in the range of 2% to 3% NaOH on pulp
in an extraction stage and are significantly lower in a
final peroxide bleaching stage. In addition, in peroxide
bleaching the pH level and the peroxide concentration
decrease during the reaction. Thus, in the upper part
of a bleaching tower the pH and the peroxide
concentration are signifi cantly lower. The optimum
parameters for hydrogen peroxide bleaching and, with
restrictions, also for hydrogen peroxide delignification
are not within the range of strong corrosion of titanium
3
metal. The corrosion of titanium is inhibited with
calcium ions (water hardness) or magnesium salts [5,
6, 7]. Already 1 ppm of calcium has a pronounced
effect [5]. Sodium silicate also inhibits corrosion [5].
Since magnesium is normally added to peroxide
bleaching processes, a corrosion of titanium under
typical mill conditions becomes very unlikely, if the
amount of chemicals added is properly controlled.
4. Prerequisites of hydrogen peroxide
application
Transition metals are decomposing hydrogen perox-
ide. A controlled decomposition with a well defined
generation of radicals would be desirable from the
point of an improvement of the delignification. Un-
fortunately untill today no selective generation has
been described. Typically the radicals produced
through metal catalysed decomposition are inselec-
tive and fiber damage dominates, as a result of
cellulose depolymerisation. In consequence, metal
impurities have to be removed from the pulp before a
subsequent peroxide treatment [8, 9, 10]. The
amount of transition metals being present in pulp
differs in a wide range and is depending on the soil
where the wood has been grown. Normally manga-
nese and iron are dominating, other metals like
copper and cobalt are present only in traces around 1
ppm. In sulphite pulping the removal of metal is very
easy. During the acidic and reducing conditions of the
pulping process, metals become water soluble and
are easily removed during brown stock washing.
In kraft pulping the transition metal ions become
insoluble during the pulping process. They are re-
duced to a low state of oxidation and precipitated as
sulphides. These sulphides are very insoluble under
alkaline and neutral conditions, so normal washing
does not remove them. During oxygen delignification
they might be oxidized to a higher state of oxidation,
but the resulting hydroxides are still insoluble under
the conditions of oxygen stage washing. They
become water soluble under mild to strong acidic
conditions. In conventional bleaching processes
transition metals are removed during the acidic
bleaching stages. Since hydrogen peroxide typically
is applied in ECF bleaching only after the first D
stage, the metal profile normally is already low
enough and no specific measures are required. Figure
2 demonstrates the effect of the pH value on the
elimination of iron and manganese from kraft pulp.
initial 6 5 4 3 2 1,4
pH value
0
20
40
60
80
100
metals (ppm)
Fe
Mn
Figure 2:
Removal of iron and manganese from softwood kraft
pulp with increasing acidity. All trials at 3% cons.,
60C, 0.5h with H
2
SO
4
for acidification.
In comparison to iron the removal of manganese
obviously is much easier. Strong acidic conditions
are required to lower the amount of iron. In TCF
bleaching the removal of the metals becomes more
important, because hydrogen peroxide is applied
early in the sequence at much higher charges.
As demonstrated in figure 3, the elimination of the
metals is improved with the addition of chelating
agents. Since compounds like EDTA are good
chelating agents for double-charged metal ions, the
addition of a small amount of sulphite improves the
chelation. DTPA can chelate also higher charged
metal ions. The metal ions present in the pulp are
very likely bound to lignin or cellulose. EDTA and
DTPA are competing with these compounds for the
metal ions. In addition, alkali earth metals are also
chelated by DTPA or EDTA, depending on their
chelation constants. Metals like iron do have chela-
tion constants, which are about two tenth powers
higher, but this is compensated by the fact that the
calcium concentration may be more than one hun-
dred times higher compared with the concentration of
iron. In kraft pulps calcium concentrations of about
5000 ppm are found. If the concentration of calcium is
so much higher compared to the concentration of
iron, then despite the higher chelation constant, the
alkali earth metals are the preferred compounds for
chelation. The demand for DTPA will therefore be
higher than the stoichiometric equivalent calculated
for iron and manganese.
The results shown in figure 2 and 3 demonstrate a
much stronger binding of iron to the residual lignin or
to the cellulose compared to manganese. Even very
high charges of DTPA do not affect the iron content at
pH 6. In contrast, manganese values are significantly
lowered. In consequence, an efficient removal of
transition metals needs higher acidic conditions, even
in the presence of a chelating agent. The necessity of
efficient washing goes without saying. Table 2 gives
an example for the metals removal from eucalyptus
kraft pulp using acidic conditions with and without
chelant addition.
4
DTPA (%): 0,25 0,5 1 all 0,25 % DTPA
initial pH 6 pH 6 pH 6 pH 8 pH 7 pH 5 pH2
0
20
40
60
80
100
metals (ppm)
Mn
Fe
Figure 3:
Removal of iron and manganese from softwood kraft
pulp with DTPA and increasing acidity. All trials at
3% cons., 50C, 0.5 h with H
2
SO
4
for acidification.
Table 2:
Removal of transition metals from oxygen delignified
eucalyptus kraft pulp. Conditions: 3% consistency,
60C, 0,5 h, 1% H
2
SO
4
, 0,5% NaHSO
3,
pH 2,3
pH
value
Fe
(ppm)
Mn
(ppm)
oxygen delignified 25 5
A
(W)
2,3 11 <1
A
(Q)
with
0,1% EDTA
2,3 10 <1
Bleaching experiments demonstrated, that it is not
necessary to decrease the metals content to ex-
tremely low levels. The addition of a small amount of
metal ions to a superchelated pulp has a more
pronounced negative effect compared with the same
amount of metal as residual. I.e. pulp chelated down
to 20 ppm iron bleaches better than pulp
superchelated down to 5 ppm iron and enriched
artificially with another 5 ppm. Very obviously the
normal residual of iron or other metals is tightly bound
to the pulp and is not available for the decomposition
reaction. In consequence there is no advantage of a
further drop of the iron level.
It is very difficult to describe a threshold value for a
"no effect level for iron or manganese, because of the
chelation of these metals to the pulp or the lignin.
Typically the residual remaining after an acid
wash/chelation treatment for iron differs between 10
and 30 ppm and for manganese between 1 and 5
ppm. If washing was efficient, the bleaching per-
formance is all right within this range of residual metal
ions.
5. Hydrogen peroxide in pulp
DELIGNIFICATION
5.1 Hydrogen peroxide in ECF sequences
As mentioned above (chapter 2), alkaline hydrogen
peroxide has a limited delignification ability. In ECF
bleaching it is applied to improve the lignin removal in
the extraction stage. In the extraction stage the
oxidized lignin is solubilized. The solubility of these
compounds in water becomes much better under
alkaline conditions, because salts are formed. The
oxidation with chlorine dioxide produces less free
phenolic groups and carboxylic acid groups on the
high molecular residual lignin compared with chlorine.
In consequence the combined effect of a chlorine
dioxide treatment and a subsequent extraction stage
is inferior compared with the result of chlorine plus
extraction. This becomes visible in figure 4.
0
2
4
6
8
10
12
14
kappa D-E resp. C-E
k-factor 0,25
k-factor 0,175
k-factor 0,1
Figure 4:
Delignification of eucalyptus kraft pulp with sequence
C-E resp. D-E and mixtures of Cl
2
and ClO
2
The addition of small amounts of oxygen to the
extraction stage is a technical possibility since high
shear mixers permit a fluidization of the pulp at
medium consistency. Oxygen gas is added to the
pulp in the form of micro bubbles, and a homogenous
mixture of pulp, gas and water is generated. In 1979
the combined addition of oxygen and hydrogen
peroxide was first introduced in a pulp mill in southern
Germany. Both chemicals oxidize lignin into better
soluble, lower molecular weight fragments. Oxygen
reacts primarily as an electrophile with the phenolic
structures, but produces also peroxo intermediates,
which cleave side chains on the residual lignin in a
nucleophile reaction.
The amount of oxygen is normally set between 5 and
10 kg per ton of pulp. It is generally accepted to
overdose oxygen, because oxygen is inexpensive.
The consumption of oxygen is a function of the
reaction temperature and the mixing efficiency.
Increasing the pressure from 2 to 3 bars to higher
levels, normally is only of benefit, if the mixer is
inefficient. If the oxygen bubbles are really small and
properly distributed, a pressure increase cannot
produce much effect. With injectors or spargers gas
bubbles of a larger diameter may result. Therefore
5
under these conditions a compression can be effec-
tive.
Because of the limited effect of hydrogen peroxide on
deligification it does not make sense to apply larger
amounts. Normally the amount of hydrogen peroxide
is kept strictly well below 10 kg per ton of pulp.
Typically about 3 to 5 kg/t are applied in an EOP
stage. Figure 5 compares the effect of an E, an EO
and an EOP treatment after chlorination and after
chlorine dioxide treatment.
C-E C-EO C-EOP D-E D-EO D-EOP
0
2
4
6
8
kappa
5
Figure 5:
Effect of oxidative reinforcement of the extraction
stage with oxygen and oxygen plus hydrogen
peroxide. Eucalyptus kraft pulp kappa 16.2; active
chlorine multiple 0.15; ~0.5 % O2, 0.5 % H
2
O
2
.
The positive effect of the addition is clearly visible. It
is smaller, if chlorine is applied and more pronounced
after the less effective delignification with chlorine
dioxide. The effect of an increasing hydrogen peroxide
charge is given in Figure 6. In the temperature range
of 70C to 80C, the effect on the kappa number
reduction is pronounced within the addition range
from zero to 0.5 % H
2
O
2
. Higher hydrogen peroxide
charges lift the brightness, but do not alter the lignin
content very much. Under these reaction conditions,
a residual of peroxide at the end of the treatment
demonstrates the absence of lignin structures which
are easily oxidized.
However, this effect can be improved in terms of
brightness increase and delignification under more
drastic conditions, e.g. higher temperature and higher
peroxide dosage. Table 3 gives a comparison of the
results achieved with increased temperature and
pressure on a D
0
pretreated pulp.
0 0,25 0,5 0,75 1
% H2O2
3
4
5
kappa
65
70
75
80
85
brightness % ISO
brightness
kappa
Figure 6:
Lignin degradation and brightness increase of
eucalyptus kraft pulp with increasing hydrogen
peroxide charge in EOP subsequent to the D0
treatment with 0.15 active chlorine multiple: EOP
conditions: 10 % cons., 75C, 30 min., 0.3 MPa O2
pressure, 60 min. without pressure.
Table 3:
Effect of temperature and pressure in an EOP
treatment of eucalyptus kraft pulp on kappa,
brightness and viscosity. D
0
with 0.15 % active
chlorine multiple. Constant conditions in EOP: 1,8%
NaOH; 10 % cons., 0.5 h pressure, 1.0 h without
pressure (simulation of upflow tube and downflow
tower).
temp.
(C)
H2O2
(%)
oxygen
(MPa)
kappa brightness
(% ISO)
viscosity
(mPa
.
s)
80 0.5 0.3 4.0 81.4 11.2
80 1.0 0.3 3.8 82.5 11.1
100 0.5 0.3 3.2 81.6 10.3
100 0.5 0,5 3.3 81.7 10.3
100 1.0 0.5 3.1 82.8 9.2
The more drastic conditions yield lower residual lignin
levels. While the effects on the Kappa number
reduction remain moderate, the increase of the
temperature basically lowers the pulp quality. At the
higher temperature level, the viscosity of the pulp
becomes lower. The improved delignification and the
possible savings of chlorine dioxide are traded for a
lower quality. In addition the higher temperature does
have a negative impact on the temperature profile of
the sequence. Normally the temperature changes
from stage to stage should not be too high, because
this might cause difficulties with the temperature level
of the washing water circuits. The increase of the
pressure does not produce a better peroxide
performance. On mill scale certainly pressure can be
important, especially if poor oxygen mixing has to be
compensated, but this has no impact on the
performance of hydrogen peroxide.
6
80 C 80 C 95 C 95 C
2
2,5
3
3,5
4
kappa
17
18
19
20
21
viscosity (mPa.s)
kappa
viscosity
MgSO4: 0,1% -- 0,1% --
Figure 7:
Effect of temperature increase and MgSO
4
addition in
EOP treatment of softwood kraft pulp. Sequence O-D
0
with kappa-factor 0,175 in D
0
. Reaction in EOP with
1,8% NaOH, 0,3% H
2
O
2
The addition of magnesium salts improves the
viscosity levels especially at higher temperature.
Figure 7 illustrates an example for softwood kraft
pulp. The increase of the EOP stage temperature
only yields a moderate drop of the lignin level from
kappa 3,4 to 3,1 (which is about 10%), and the
viscosity losses can kept moderate, if magesium
salts are added.
With the addition of oxygen and peroxide in the
extraction stage a decrease of the residual lignin level
down to kappa numbers between 3 and 4 can be
achieved with an active chlorine multiple in the D
0
stage as low as 0.15 to 0.2. Temperatures above the
usual level of 75C to 85C in an EOP stage cannot
be recommended since the benefit in terms of de-
lignification is moderate and the pulp quality suffers.
In ECF bleaching the addition of hydrogen peroxide to
the first extraction stage has these advantages:
Y H
2
O
2
improves the delignification and increases
the brightness
Y small amounts of H
2
O
2
already yield a pronounced
effect
Y H
2
O
2
addition can be combined with oxygen
Y standard E stage temperature produces good
results.
5.2 Hydrogen peroxide in TCF sequences
In TCF bleaching sequences delignification is the
biggest problem. Basically only three bleaching
chemicals are available to decrease and bleach the
lignin residual. Oxygen cannot oxidize lignin as
efficiently as chlorine. Typically the oxygen stage is
run at maximum effect, but the residual lignin level
remains high. Ozone is very effective in lignin oxi-
dation, but also reacts with cellulose. In addition it is
expensive and difficult to mix in higher amounts in the
medium consistency range prefered in pulp mills.
Thus high ozone charges cannot be applied. This
leaves a part of the delignification work for hydrogen
peroxide. As mentioned above, hydrogen peroxide is
not an ideal delignification agent. It needs a con-
siderable amount of pushing, if more than just a small
margin of delignification has to be achieved.
TCF pulp bleaching technology started with the
application of only two bleaching chemicals. This
restricted the brightness gain to the semibleached
level [8]. Aspa Bruk mills almost classical approach
for TCF bleaching is basically a very simple one:
1. Push oxygen delignification to its limits,
2. intensify the oxidative extraction stage with a high
peroxide charge,
3. use high temperature in combination with a lot of
peroxide and extent the reaction time as long as
the tower capacity permits.
As any simple concept it has the advantage to work.
On the other hand it does not yield miracles. Bright-
ness is a function of the kappa number and the
delignification with this concept is limited. The first
TCF pulps reached only 72 to 75 % ISO brightness.
It is difficult to push delignification with oxygen by just
repeating the addition of oxygen. Two stages of
oxygen treatment indeed improve the lignin removal,
but the effect remains moderate [11]. Unfortunately a
conventional oxygen stage is run with a very high
carry-over of organic compounds, due to the
countercurrent water flow into the recovery system.
This sets limits to the performance of the oxygen
treatment, because the organic material as well as
the metal ions promote side reactions and are
prohibitive for the application of a higher temperature.
For softwood kraft pulp a total delignifi cation of about
60% is the highest level that can be reached without
severe cellulose degradation. The remaining amount
of lignin is still too high to allow very high brightness
levels in a final treatment with peroxide only.
An alternative to increase the delignification is the
addition of hydrogen peroxide. The prerequisite for the
application of hydrogen peroxide is the elimination of
the transition metals which remain in the pulp after
conventional oxygen delignification. An acidic
washing stage removes the metal ions. This washing
stage has another positive side effect, the pulp enters
the oxygen/peroxide stage without a high load of
organic material. The temperature in the oxygen
treatment can be increased without severe viscosity
losses. This is an advantage also for the consumption
of hydrogen peroxide, which would otherwise not
react completely. The effect of the addition of
7
hydrogen peroxide to a second oxygen stage
following an acidic washing stage is described in
figure 8. The dominant amount of additional de-
lignification is achieved already with the second
addition of oxygen at the higher temperature. The
effect on pulp viscosity remains moderate. The
addition of 1% or 2% H
2
O
2
increases the brightness
of the pulp significantly and allows an additional small
drop of the lignin level.
first Ostage
H2O2 charge (%): -- 1 2 -- 2
15,8
10,5
8,3
7,7
9,5
6,4
100C 100C 100C 110C 110C
5
7
9
11
13
15
17
kappa
Figure 8:
Delignification of softwood kraft pulp with a second
oxygen treatment with and without hydrogen peroxide
addition. Effect of temperature and hydrogen peroxide
addition on Kappa number. First oxygen treatment to
kappa 15,8 and 26,8 mPa
.
s
The drawback of the addition of high peroxide charges
is not only cost increase. Side reactions are taking
place, which become visible as viscosity losses and
a higher COD load, which corresponds to a yield
drop. Figure 9 shows the resulting data for viscosity
and the effluent load. The moderate drop of the
viscosity with the second oxygen treatment is
intensified by the addition of peroxide. Very obviously
side reactions, like the thermal cleavage of peroxide
into hydroxyl radicals, result in an attack of the
cellulose chains. Considering the small additional
kappa number drop, the steep increase of the COD
load can be attributed more to the degradation of
cellulose than to lignin removal.
first Ostage
H2O2 charge (%): -- 1 2 -- 2
100C 100C 100C 110C 110C
10
15
20
25
30
viscosity (mPa.s)
15
20
25
30
35
COD (kg/t)
viscosity
COD
Figure 9:
Delignification of softwood kraft pulp with a second
oxygen treatment with and without hydrogen peroxide
addition. Effect of temperature and hydrogen peroxide
addition on viscosity and effluent load. First oxygen
treatment to kappa 15,8 and 26,8 mPa
.
s
For hardwood kraft pulp the results are very similar.
Table 4 has the data for the treatment of an oxygen
predelignified kraft pulp. Following the metals
elimination, the second oxygen treatment decreases
the residual lignin level from kappa number 12
depending on the temperature to kappa values of 8,4
resp. 6,5. The addition of hydrogen peroxide yields a
significant increase in brightness, but only a
moderate effect on the lignin. The lower lignin residual
is accompanied by a loss of viscosity. Whether this
brightness increase in the EOP stage is of benefit for
the final brightness level, will be answered in chapter
6.2, together with the question of the total demand for
hydrogen peroxide to reach high brightness levels.
Table 4:
Activation of a second oxygen stage with temperature
and hydrogen peroxide addition. Eucalyptus kraft
pulp, kappa 12 after first oxygen stage, acid wash
with 0,1% EDTA; constant: 2% NaOH, 0,1% MgSO
4
,
10% cons., 1,5 h, 0,3 MPa O
2
tem-
perature
(C)
kappa bright-
ness
(% ISO)
viscosity
(mPa.s)
O 90 8,4 58,2 22,1
O 100 7,4 62,3 21,8
O 110 6,5 64,8 21,7
OP 110 6,0 82,2 17,1
peroxide addition in OP: 2% H
2
O
2
In TCF bleaching sequences hydrogen peroxide can
be applied to decrease the residual of lignin in a
second oxygen stage, which is normally a modified
extraction stage. It is important to remove transition
metals before hydrogen peroxide is added. If the
oxygen addition is efficient, i.e. a high shear mixer is
applied, oxygen itself is already very effective. This
decreases the potential benefit of the peroxide
addition:
Y Hydrogen peroxide addition to a second oxygen
stage (EOP stage) yields a high brightness in-
crease -
Y the effect on delignification remains moderate -
Y at high temperature viscosity is negatively affected
and the effluent load increases.
8
6. Hydrogen peroxide in pulp BRIGHTEN-
ING
6.1.Hydrogen peroxide in ECF sequences
Hydrogen peroxide is applied in the final stages of an
ECF bleaching sequence with the target to optimize
the requirement for bleaching chemicals. Basically in
each extraction stage of a bleaching sequence the
addition of hydrogen peroxide is possible. The
question is, whether such an addition offers an
economical advantage. For northern softwood and the
bleaching sequence CD-E
1
-D
1
-E
2
-D
2
this question was
positively answered already in 1981[12]. The addition
of a small amount of hydrogen peroxide in the E
2
stage was economical, because it lowered
significantly the chlorine dioxide requirement for final
bleaching.
Another variation of the addition of hydrogen peroxide
became very typical in the United States: high
density storage addition of hydrogen peroxide. It
allowed pulp mills to boost the brightness in an ad-
ditional bleaching stage without spending capital. In
the race for top market pulp brightness and the
resulting top selling price advantage, the additional
two to three brightness points which are easily
achieved with as little as 2 kg to 3 kg of H
2
O
2
do pay
off. Certainly high density storage application of
hydrogen peroxide is only a compromise. Neverthe-
less, it makes sense, because the capital require-
ment for an additional bleaching stage is high, and
despite the always changing conditions in a storage
tower, the results are rather satisfying. The good
results of high density storage bleaching open an-
other alternative: a final bleaching stage run with
hydrogen peroxide.
The application of hydrogen peroxide in a final
bleaching stage is an answer to the concept of
shortening the bleaching sequences and also the
answer to the elimination of the second extraction
stage. Instead of a D
1
-E
2
-D
2
sequence the D
1
-D
2
alternative was created to cut investment costs. It
certainly needs more active chlorine compared with
the longer sequence, but the savings in capital for
one tower and one washer were believed to pay for.
On the other hand, in some mills the higher chlorine
dioxide requirement causes problems because of the
limitation in the installed chlorine dioxide capacity.
An increasing pulp production and the implementation
of ECF bleaching can also tighten the availability of
chlorine dioxide.
Bleaching an eucalyptus kraft pulp with the sequence
O-D
0
-EOP-D
1
-D
2
requires about 3,3 kg of active
chlorine, to produce the market pulp target brightness
of 90
+
ISO. The lowest demand results with an active
chlorine multiple of 0,15 to 0,175 in D
0
. The combined
active chlorine multiple for the D
1
-D
2
stages is
relatively high, a kappa factor of 0,5 is needed to
achieve the target brightness. The typical increase of
the brightness during the D
1
and D
2
stages is
demonstrated in figure 10.
0 0,25 0,5 0,75 1 1,25 1,5 1,75 2
active chlorine (%)
87
88
89
90
91
brightness (%ISO)
D1
D2
P
Figure 10:
Brightness increase of an O-D
0
-EOP prebleached
eucalyptus kraft pulp (kappa number 3.5 and 82.1 %
ISO brightness) with ClO
2
or H
2
O
2
applied in final
bleaching at 10 % cons. D stages 75C, 3 h, P
stage: 0.25 % H
2
O
2
, O.4 % NaOH, 80C, 3 h.
In the D
1
stage increasing amounts of chlorine dioxide
yield a significant brightening effect. Nevertheless,
this effect levels of at higher chlorine dioxide charges.
After an inital steep increase of the brightness, a
doubling of the ClO2 charge does not significantly
further lift the brightness. With only 0.5 % active
chlorine the brightness is lifted from 82,1% ISO to
87.8 %, with twice the amount of chlorine dioxide (1
% active chlorine), only 88.4 % ISO result. The curve
for the brightness development in D1 becomes very
flat, thus even very high charges of ClO2 do not
produce brightness levels above 89 % ISO. This is
the reason why indeed an additional stage (D
2
) is
required and in the past a D
1
-E
2
-D
2
treatment was
applied. The desired final brightness above 90 % ISO
is achieved only after washing and the addition of
more chlorine dioxide in the second D stage. Under
these conditions a total input of 1.5 % active chlorine
yields 90 % ISO brightness.
Changing the bleaching conditions, for example the
pH environment from acidic to alkaline, or applying a
different bleaching agent, normally results in an
advantage compared with the repeated addition of the
same chemical. So, instead of two subsequent
chlorine dioxide treatments, a chlorine dioxide stage
followed by an alkaline peroxide stage should be
more efficient. This becomes obvious in the left part of
figure 10. Very low levels of chlorine dioxide in D
1
yield a sufficient activation to achieve the required
brightness increase in the peroxide stage. In
consequence, a significantly lower demand for
chlorine dioxide is needed to achieve the brightness
target. For another pulp sample the results are very
similar, as is demonstrated in figure 11. Mill scale
trials have reproduced these results and demon-
strated its general applicability [13]. This effect can
9
be very attractive to mills with a limited chlorine
dioxide capacity, or for mills which want to increase
pulp production but not their chlorine dioxide gen-
erator capacity.
90,7
91,3 91,3
90,4
91
90,9
88,7 88,7
89,2 89,2
89,3 89,3
0,1 0,1 0,15 0,15 0,2 0,2
active chlorine multiple in D1
88
89
90
91
92
brightness (% ISO)
brightness D1
0,1% H2O2
0,2% H2O2
Figure 11:
Brightness increase for different kappa-factors in D
1
and peroxide charges. Pre-bleaching with sequence
O-D
0
-EOP-D
1
with oxygen stage kappa 11,1; kappa-
factor 0,15 in D
0
; P stage with 0,1 % MgSO
4
, 75C,
1,5 h 10% consistency
The amount of hydrogen peroxide needed to boost
the brightness is rather low. Figure 12 demonstrates
the brightness increase achieved with an increasing
hydrogen peroxide charge. Very small amounts of
hydrogen peroxide already produce a brightness
increase of about two points. In addition the bright-
ness stability is improved. A comparison of the re-
sults achieved with a final D
1
-D
2
and a D
1
-P treatment
demonstrated a decrease from the average 4 to 5
points brightness loss after D
1
-D
2
to only 2 to 3 points
after D
1
-P. In consequence, the target brightness on
mill scale can be decreased. A calulation of the
savings in chlorine dioxide against the demand for
hydrogen peroxide shows a replacement ratio of up to
9 kg of chlorine dioxide (as active chlorine) with only
1 kg of hydrogen peroxide. The active chlorine
multiple for the D
1
-D
2
stages can be lowered from 0,5
to a level as low as 0,1 for the D
1
-P combination.
0 0,05 0,1 0,15 0,2 0,25
H2O2 charge (%)
87
88
89
90
91
brightness (%ISO)
Figure 12:
Effect of an increased hydrogen peroxide charge in
postbleaching. Bleaching of an O-D
0
-EOP-D
1
prebleached pulp. Kappa-factor in D
0
0.17, in D
1
0.2.
The viscosity level in the final peroxide stage can be
controlled with the addition of magnesium salts. As
expected, the temperature of the P stage has an
effect on pulp viscosity. A high peroxide residual also
has a negative impact on the viscosity, especially at
higher temperature and neutral pH. Table 5 gives the
results for a variation of the bleaching conditions in
the final peroxide stage. The much better thermal
stability of the peroxide bleached pulps becomes
obvious, if heat aging brightness losses are compared
with standard D
1
-D
2
bleached pulp.
In a final peroxide stage the application of pressure
was recommended [14]. It is very difficult to
understand why this recommendation was made. A
small amount of hydrogen peroxide reacts already at
80C within two hours without the application of
pressure. Figure 11 illustrates this. At higher
temperature the viscosity values suffer and the
brightness is not tremendously improved.
Table 5: Improvement of the final viscosity after peroxide bleaching of eucalyptus kraft pulp. Pulp prebleached
with O-D
0
-EOP-D
1
to 87.7 % ISO brightness. Constant: 10 % cons., 2 h.
chemicals addition (%) temp. (C) residual
H
2
O
2
bright-
ness
viscosity
(mPa.s)
heat
ageing
H2O2 NaOH MgSO4 sodium
silicate
(%) (% ISO) (points)
0.25
0.25
0.50
-
0.4
0.4
-
-
-
-
-
-
70
70
70
0.02
0.01
0.10
87.6
90.2
90.8
8.5
10.8
10.1
7.2
2.0
2.8
0.25
0.25
0.25
0.4
0.3
0.3
0.1
0.1
0.1
-
0.25
0.50
70
70
70
0.04
0.04
0.05
90.5
90.7
90.8
11.2
13.4
13.8
2.6
2.6
2.4
0.1 0.3 0.1 0.50 70 0.03 90.5 13.9 2.4
0.1
0.2
0.3
0.3
0.1
0.1
0.50
0.50
90
90
0.01
0.01
90.7
90.9
12.8
12.5
2.5
2.6
control (D1D2 treatment): 75 - 90.4 13.2 4.6
10
In consequence for the application of hydrogen
peroxide in the final stage of ECF sequences our
recommendation is not to decide for drastic
conditions, but for a smooth reaction at moderate
conditions. A final P stage offers these advantages:
Y Small amounts of hydrogen peroxide yield high
brightness gains already at conventional
temperature levels -
Y the brightness stability is significantly inproved.
The selection of the best conditions in the P stage has
to be made with following facts:
Y magnesium salt addition increases viscosity -
Y buffering yields maximum brightness at highest
viscosity -
Y high temperature increases the brightness but
yields lower viscosity -
Y viscosity is negatively affected by residual peroxide
at neutral pH.
6.2 Hydrogen peroxide in TCF sequences
The basic approach to TCF bleaching with hydrogen
peroxide is the simple translation of the available
experience with mechanical pulp bleaching to
chemical pulp. Metals elimination under mild acidic
conditions in the presence of chelating agents or just
washing under more acidic conditions generates the
basis for peroxide bleaching. High consistency and
buffered conditions improve the performance of hy-
drogen peroxide and push the reaction to develop
brightness and not dominantly lignin degradation.
Certainly, under these conditions delignification takes
place too, but not with the same intensity. This is
demonstrated in figure 13, using a spruce sulphite pulp
as an example. Under buffered conditions more
brightness is generated and the decrease of the kappa
number is smaller.
Na- silicate (%) none 4 none 4
unblea. 1% 1% 2% 2%
0
5
10
15
20
kappa
55
60
65
70
75
brightness (%ISO)
kappa
brightness
H2O2
Figure 13:
Simultaneous brightening and delignification of spruce
sulphite pulp with hydrogen peroxide with caustic soda
or caustic soda and sodium silicate. Conditions: 70C,
1,5 h, 10% consistency
Semibleaching of any type of pulp with hydrogen
peroxide yields very good results at high consistency.
The higher relative concentration of the bleaching
chemicals and the elimination of dissolved organic
compounds is favorable for the performance of the
bleaching process [15]. European sulphite mills have
modified their bleach plants to a combination of a
medium consistency delignification stage and a high
consistency bleaching stage [16]. Under high
consistency conditions the activation of the peroxide
reaction with alkali is achieved with very small
amounts of caustic soda, see Table 6. A partial
substitution of caustic soda with soda ash (Na
2
CO
3
) is
possible. This produces a very moderate effluent load
(COD) which can be translated into a very low yield
loss of the bleaching process [15].
Table 6:
Decrease of effluent load with high consistency hy-
drogen peroxide bleaching. Spruce sulphite pulp
predelignified with EOP to kappa 7,6; brightness
75,3% ISO, constant peroxide charge 2%.
consis-
tency
(%)
NaOH
(%)
Na
2
CO
3
(%)
Na-
silicate
(%)
bright-
ness
(% ISO)
COD
(kg/t)
10 1,6 - - 85,0 29
30 0,5 - 0,75 86,7 17
30 0,25 0,25 0,75 87,1 15
Similarly kraft pulps can be bleached with hydrogen
peroxide. Because of the more condensated residual
lignin, the prerequisite for the bleaching of kraft pulps
is a sufficient amount of delignification in the first
bleaching stages [8]. The bleachability of a kraft pulp
is depending on the pulping conditions and the
intensity of the delignification conditions in the first
bleaching stages. Typically hardwood pulps bleach
much easier compared with softwoods. The degree of
delignification is an important factor for the final
brightness level.
The basic delignification in TCF sequences is typically
achieved with one or two oxygen stages. To improve
the effect, a treatment with ozone or peracid can be
conducted between the oxygen stages. An alternative
is an ozone stage following the delignifi cation with
oxygen. The second oxygen stage can be activated
with hydrogen peroxide as described in chapter 5.2.
Three main problems exist in TCF bleaching:
1. to achieve a kappa number low enough to produce
the required brightness reliable,
11
2. to achieve this brightness with an acceptable input
of bleaching chemicals,
3. to achieve an acceptable pulp strength.
The first TCF softwood kraft pulps available on the
market were bleached with a combination of oxygen
delignifcation, metals removal, intensified oxidative
extraction and a subsequent peroxide treatment. This
was achieved with rather high hydrogen peroxide
charges. As mentioned above, this procedure is the
translation of the existing experience with mechanical
pulp to kraft pulp and consequently the European
Patent Office denied to grant a patent. Because the
resulting kappa number was still high, only brightness
levels of 70 % ISO to 75 % ISO were gained. Modified
pulping methods, intensified oxygen delignification and
increasing temperature levels in the peroxide stage
resulted in a continuous improvement of the brightness
to about 80 % ISO [17, 18]. The introduction of ozone
into the bleaching sequences allowed a further
increase of the brightness level. Still brightness 85 %
ISO and more require a rather high input of hydrogen
peroxide. Published peroxide requirements are about 3
% H
2
O
2
[19], but in actual mill practice 4 % and even
5% of hydrogen peroxide are applied. This is certainly
also the result of inefficient mixing of the applied ozone
charge and of not optimum conditions in bleaching. It
looks as if medium consistency conditions only allow
the addition of very small amounts of ozone.
Peracids are an alternative to ozone for the activation
of pulp prior to a second oxygen stage. If ozone is
applied under high consistency conditions, very little
ozone is required to yield a pronounced effect. In
comparison to ozone the effect on delignification with
peracids is limited. The advantages of peracids are the
possibility of using standard medium consistency
equipement, the option of maintaining the temperature
level and the potentially simpler on-site generation. A
comparison of the results of the activation with peracid
or ozone shows a clear advantage for ozone in terms
of delignification (Table 7). Very small amounts of
ozone give a significant drop of the kappa number after
the subsequent oxygen stage [20]. This kappa number
decrease is accompanied by a drop of the viscosity
values. Peracids do produce less delignification, but
do have only a very moderate effect on the viscosity.
Table 7:
Activation of a second oxygen treatment with peracid
or ozone. Pulp pretreated with sequence O-A
(Q)
- to
kappa 10; ozonation at 30% cons., pH 4 to 5, ambient
temp.; peracid stage at 10% consistency, 70C, 1 h,
pH
end
4,2 to 3,5; second oxygen stage with 2% NaOH,
0,1% MgSO
4
, 0,3 MPa O
2
at 110C, 1,5 h, 10%
consistency
sequence per-
acetic
Caro
s acid
ozone oxygen stage
results
acid
(%) (%) (%) kappa
#
vis-
cosity
(mPas)
O-A
(Q)
-O - - - 6,5 21,7
O-A
(Q)
-Paa-O 0,75 - - 5,5 21,6
O-A
(Q)
-Paa-O 1 - - 4,7 21,4
O-A
(Q)
-Paa-O 1,5 - - 4,1 20,6
O-A
(Q)
-Ca-O - 1,5 4,4 19,1
O-A
(Q)
-Z-O - - 0,2 5,8 18,8
O-A
(Q)
-Z-O - - 0,3 4,9 17,1
O-A
(Q)
-Z-O - - 0,5 2,3 12,5
In order to select the most appropriate pretreatment,
the bleachability of these pulps has to be compared. If
it would be unnecessary to delignify to very low kappa
numbers, if pulps with a higher residual lignin level
could be bleached to a high final brightness. The
advantage of the still high viscosity at a higher kappa
level after the second oxygen stage could be utilized.
Thus the question is, whether kappa numbers around
4 already allow to achieve brightness levels above 90
% ISO with an acceptable peroxide input.
In bleaching experiments with a pulp with a relative
high residual of lignin ( kappa 6) the benefit of the
addition of magnesium sulphate and sodium silicate
became visible (Table 8). Brightness and viscosity
were highest in the presence of both chemicals. In the
absence of the stabilizers, the consumption of
hydrogen peroxide was higher, but brightness and
viscosity were low. Thus, a high turnover of peroxide
not necessarily is an advantage. Despite the high
reaction temperature of 90C and the long retention
time of 8 hours in all experiments the amount of
unconsumed hydrogen peroxide was high. The op-
timization of the peroxide consumption with increased
temperature and extended retention time is sumarized
in Table 9.
Table 8:
Stabilization of hydrogen peroxide bleaching with
sodium silicate and magnesium sulphate. Bleaching of
kappa 6 eucalyptus kraft pulp with 4% H
2
O
2
at 90C,
10% cons., 8 h.
trial NaOH
(%)
Na-
silicate
(%)
MgSO
4
(%)
H
2
O
2
residual
(%)
bright-
ness
(%ISO)
visco-
sity
mPas
1 1,5 - - 1,22 85,8 11,2
2 1,4 0,5 - 1,71 86,2 14,4
3 1,4 1 - 1,72 86,4 14,6
4 1,5 - 0,1 1,68 86,8 12,3
5 1,5 - 0,2 1,72 86,9 12,5
12
6 1,4 0,5 0,1 1,88 87,5 14,7
Table 9:
Peroxide bleaching at 75C, 90C and 110C; with 4%
H
2
O
2
, 1,5% NaOH, 0,5% sodium silicate, 0,1%
MgSO
4
, Pulp predelignifed with the sequence O-A
(Q)
-
Ca-O-A
(Q)
to kappa 4,0.
trial tempe-
rature
(C)
consis
tency
(%)
time
(h)
H
2
O
2
residual
(%)
bright-
ness
(%ISO)
1 75 25 24 0,91 89,3
48 0,37 90,5
2 90 25 4 1,07 89,1
6 0,45 90,2
3 90 10 8 1,67 89,3
24 0,74 90,8
4 110 10 2 0,68 88,1
3 0,24 88,5
Chemical reactions are accelerated with a higher
concentration of the reaction partners and by a higher
reaction temperature. The application of high
consistency increases the relative concentration of
bleaching chemicals to pulp. At 75C this produced
the desired top brightness of more than 90% ISO, but
required two days of reaction. An increase of the
temperature to 90C did cut the reaction time to only 6
hours. Nevertheless, high consistency is not prefered
in kraft pulp bleach plants, therefore the other
alternative, temperature increase was also tested. The
highest brightness was produced after 24 hours
reaction time at 90C, the increase of the temperature
above the boiling point of water to 110C yielded a very
fast consumption of the hydrogen peroxide, but not an
equivalent brightness increase. Obviously the high
reaction temperature favors peroxide consumption, but
not a consumption which yields bleaching reaction,
but consumption based on side reactions.
Therefore the comparison for the bleachability following
ozone or peracid activation was conducted with only
90C as reaction temperature and one day of reaction
time. The results in table 10 demonstrate a better
bleachability with peroxide following the ozone/oxygen
treatment. Very little ozone is required to achieve a
brightness above 90% ISO with only 2% of H
2
O
2
. The
advantage of the higher viscosity after the peracid
treatment is more than compensated by the better
increase of the brightness at a higher residual of lignin.
With ozone less chemical yield better delignification
and on top of this, brightness increase is higher with
less peroxide input at the higher lignin residual.
Because high charges of ozone are not required, the
final viscosity values become similar to those with the
peracid pretreatment.
Table 10:
Comparison of the sequences O-A
(Q)
-Z-O-A-P and O-
A
(Q)
-Paa-O-A-P. Peroxide bleaching at 90C, 10%
consistency, 24h
per-
acetic
acid
(%)
ozone
(%)
kappa
after
oxygen
H
2
O
2
(%)
bright-
ness
(%ISO)
visco-
sity
(mPa
s)
0,3 4,9 2 90,7 14,8
0,3 4,9 2,5 91,8 14,7
0,5 2,3 2 90,9 13,2
0,5 2,3 3 91,8 11,0
0,75 5,5 3 88,4 15,5
1,5 4,1 2 88,8 19,1
1,5 4,1 3 90,5 15,0
These results are satisfying in terms of viscosity,
chemicals requirement and brightness. Brightness
increase is best after ozone activation and very little
ozone is needed to produce a brightness above 90 %
ISO in the final peroxide stage. There is one open
question from chapter 5.2 remaining to be answered,
the question whether a prebleaching with hydrogen
peroxide in the second oxygen stage to improve
delignification also produces a good bleachability in
the final bleaching stage. Table 11 gives an answer.
The significant brightness increase achieved with the
addition of hydrogen peroxide to the second extraction
does not have a pronounced effect on the demand for
hydrogen peroxide in the final bleaching stage. Despite
the high brightness after the OP treatment (82,2 %
ISO, see table 4), the final peroxide stage still requires
a high peroxide amount to produce 90 % ISO.
Compared with the addition of peroxide to the second
O stage, ozone activation is much more effective.
Table 11:
Comparison of the sequence O-A
(Q)
-Z-O-A-P and O
A
(Q)
-OP-A-P. Bleaching in P at 90C, 10 %
consistency, 24h
sequence ozon
e
(%)
H
2
O
2
in 2
nd
O
(%)
kappa
after
oxyge
n
H
2
O
2
in P
(%)
bright-
ness
%ISO
viscosit
y
(mPas)
O-A
(Q)
-Z-O-
A-P
0,3 -- 4,9 2 90,7 14,8
O A
(Q)
-OP-A-
P
-- 2 6,0 2 89,8 13,2
13
Still the time required to achieve the good brightness
levels of table 9 is extremely long. A retention time of
24 hours would require an enormous volume for the
bleaching tower. The alternatives are either a high
consistency treatment as described in table 9, or a
medium consistency treatment at a temperature
somewhere between 90C and 110C.
2 4 6 8 12 16 20 24
time (h)
88
89
90
91
92
brightness (% ISO)
1,25% NaOH
1,50%
1,75%
2,00%
2,25%
Figure 14:
Effect of increasing caustic soda charge on brightness
increase with time, bleaching with 3% H
2
O
2
, 0,5 %
sodium silicate, 0,1% MgSO
4
, at 15% cons., 95C.
Another option is the acceleration of the reaction with
higher charges of caustic soda. This is demonstrated
in figure 14 for a peroxide charge of 3% at 95C. The
brightness development with time at different caustic
soda charges indeed is faster with higher alkali
amounts. Unfortunately, the faster increase does not
yield a higher brightness, the reaction only levels off
earlier. While a caustic soda charge of 1,25% does
not produce sufficient activation to produce very high
brightness even with very long retention time, alkali
charges of 1,75% or 2% NaOH yield a rapid
comsumption of the hydrogen peroxide, but only a
limited brightness increase. In consequence, it does
not make sense to apply very high charges of caustic.
The reason is the faster consumption of the peroxide
in side reactions. The decrease of the H
2
O
2
amount
with time is shown in figure 15.
A comparison of brightness achived and peroxide
consumed clearly teaches the advantage of the lower
caustic soda charges. The development of the highest
brightness still needs a full day of retention time, but
with the addition of 1,5% of caustic soda, after 12 to
16 hours, the brightness is very close to the optimum,
only about one point short of the highest value. The
negative side effect of high caustic soda charges is the
increasing extraction effect. The addition of more
caustic soda therefore also results in a lower yield.
Especially at higher temperature this effect becomes
very pronounced and puts some question marks on
the recomendation of such harsh conditions in the
literature [21]. A reaction temperature above 100C
requires pressurized conditions. Recently it was
recommended to pressurize the peroxide treatment in
order to achieve a better brightness increase in TCF
bleaching [22]. Unfortunately no mill data with
pressurized conditions are available, the conditions
cited in [22] were not pressurized. In figure 16 the
COD values produced with different caustic soda
charges at 95C and 110C are compared. The higher
COD load which results from the high temperature
treatment is clearly visible.
H
2
O
2
residual (%)
2 4 6 8 12 16 20 24
time (h)
0,4
0,6
0,8
1
1,2
1,4
1,6
1,25% NaOH
1,50%
1,75%
2,00%
2,25%
1,5 % NaOH, 110C
Figure 15:
Effect of increasing caustic soda charge on peroxide
consumption with time, bleaching with 3% H
2
O
2
, 0,5 %
sodium silicate, 0,1% MgSO
4
, at 15% cons., 95C.
1,25 1,5 1,75 2 2,25
NaOH (%)
6
8
10
12
14
16
18
COD (kg/t)
95C
110C
Figure 16:
Effect of increasing caustic soda charge on the COD
load of hydrogen peroxide bleaching at different
temperature, conditions see figure 14
According to the rules of thermodynamics, the appli-
cation of a very high pressure very likely hinders the
decomposition of hydrogen peroxide into oxygen and
water (equ. 2). The pressure needed is very high, it
cannot be expected that only some atmospheres are
sufficient. Experiments in a laboratory scale high
shear mixer which permitted the mixing of hydrogen
peroxide in the presence of oxygen at temperature
with the preheated pulp, did not show any benefit of
pressure in the range of one to four atmospheres [20].
Within the range of reproducability, the results shown
in table 12 were identical. It is difficult to answer the
14
question, why different results with positive effects are
reported in the literature [23], it looks as if the results
are depending on the laboratory scale conditions and
the equipement used for the trials.
Table 12:
Effect of pressurizing a peroxide bleaching stage with
oxygen at 110C, pre- treatment with O-A
(Q)
-EOP- to
kappa 6,0; bleaching with 4% H
2
O
2
, 1,5% NaOH, 0,5%
sodium silicate, 0,1% MgSO
4
, at 10% cons., 2h in a
high-shear mixer.
trial oxygen
pressure
MPa
H
2
O
2
residual
%
brightness
%ISO
viscosity
mPas
1 - 0,76 86,4 15,8
2 0,2 0,81 86,6 15,4
3 0,3 0,79 86,3 15,1
4 0,4 0,82 86,7 15,3
Table 13 summarizes the compromise for the best
results in terms of retention time and brightness
increase. A total time demand of 8 hours retention
time at 100C seems to be acceptable, it results in a
brightness above 90% ISO, which is higher compared
with the very fast, but less efficient reaction at 110C.
Table 13:
Optimization of the retention time/temperature
relationship in peroxide bleaching. Pulp predelignified
to kappa 4,3 with the sequence O-A
(Q)
-Z-O-A; const.
2,5% H
2
O
2
, 0,5% sodium silicate, 0,1% MgSO
4
tempe-
rature
(C)
consi-
stency
(%)
NaOH
(%)
time
(h)
bright-
ness
% ISO
visco-
sity
mPas
95 15 1,5 8 90,0 13,9
16 91,4 13,7
100 15 1,75 8 90,3 13,8
12 90,5 13,6
110 10 1,5 2 88,5 13,3
The recommendation for the application of hydrogen
peroxide in TCF sequences therefore is to concentrate
the hydrogen peroxide input on one or two final
bleaching stages. The prerequisite is a sufficient
delignification with oxygen and ozone is the earlier
bleaching stages. The performance of high hydrogen
peroxide charges is best at temperatures lower than
100C because decomposition and intensified ex-
traction are the negative side effects of higher tem-
perature. The conditions for the peroxide stage are
very flexible, but in order to have an acceptable re-
tention time, the temperature and the consistency
have to be high. Other parameters of importance are:
Y magnesium salt addition increases viscosity -
Y buffering yields maximum brightness at highest
viscosity -
Y high temperature increases the brightness, but with
very high temperature bleaching becomes
inefficient.
7. Summary
The final conclusions in this look over peroxide ap-
plication and performance in chemical pulp bleaching
can be made as follows:
Hydrogen peroxide is a very versatile bleaching agent.
It can be applied in a wide range of conditions:
Y Temperature range: (starting at about 20C)
normally: 60C to 75C,
in special cases up to 100C.
Y Retention time range: (starting with about 30
min.)
normally: 1,5 h to 2 h and up to 16 h,
in special cases 1 day to 3 days.
Y Consistency range: not below 8 % consistency,
normally: 10% to 15% consistency,
in special cases up to 30% consistency.
Y In ECF bleaching sequences the application of
hydrogen peroxide offers significant economical
benefits. The addition of hydrogen peroxide in the
extraction stage improves the delignification and
allows lower kappa factors in D
0
.
Y Final bleaching with hydrogen peroxide cuts the
requirement for chlorine dioxide in D
1
. A very low
input of hydrogen peroxide allows kappa factors in
D
1
as low as 0,1 to 0,15.
Y The decreased demand for chlorine dioxide yields
lower cost, lower AOX of the effluent and a lower
OX in the pulp.
Y The moderate conditions (pH and concentration)
required for bleaching with hydrogen peroxide do
not result in a corrosion hazard for titanium equi-
pement.
Y In TCF bleaching sequences hydrogen peroxide
is the chemical of choice to produce the required
brightness level.
Y Peroxide bleaching yields highest brightness in-
crease at temperature levels between 70C and
90C with long and very long retention time.
Y For kraft pulps the best compromise is an increase
of the temperature close to the boiling point.
YAt high temperature the addition of magnesium salts
and buffering with very small amounts of soduim
silicate is the best strategy to obtain highest
viscosity values.
15
8. References
[1] N. N. Greenwood and A. Earnshaw; Chemistry of
the Elements; Pergamon Press, Oxford (1984)
[2] J. Takagi and K. Ishigure; Thermal Decompo-
sition of Hydrogen Peroxide and its Effect on
Reactor Water Monitoring of Boiling Water Re-
actors; Nuclear Science and Engineering: 89,
177 - 186 (1985)
[3] J. Gierer; The Chemistry of Delignification;
Holzforschung 36, (2) 55 - 64 (1982)
[4] G. Gellersted and I. Petterson; Chemical As-
pects of Hydrogen Peroxide Bleaching. Part II:
The Bleaching of Kraft Pulp; J. Wood Chem. and
Techn. 2 (3) 231 - 250 (1982)
[5] R. W. Schutz and M. Xiao; Development of
practical guidelines for titanium in alkaline per-
oxide bleach solutions; Intn. Pulp Bleaching
Conf. Poster Proceedings 153 - 157 (1994)
[6] S. J. Clarke and D. L. Singbeil; Corrosion of
titanium in alkaline peroxide bleaching media;
Pulp Paper Canada 95 (10) T417 - T421 (1994)
[7] J. MacDiarmid, R. Charlton and D. L. Reichert,
Corrosion and material engineering considera-
tions in hydrogen peroxide bleaching; TAPPI 7th
Intn. Symposium on Corrosion Proceedings, 97 -
104 (1992)
[8] H. U. Sss, W. Eul, O. Helmling; Semibleaching
of kraft pulp using oxygen and hydrogen
peroxide; Papier 43 (7) 318 - 323 (1989)
[9] J. Bouchard, H. M. Nugent and R. M. Berry; A
comparison between acid treatment and chela-
tion prior to hydrogen peroxide bleaching of kraft
pulps, J. Pulp and Paper Science; 21, (6), J203 -
J208 (1995)
[10] L. Lapierre, J. Bouchard, R. M. Berry and B. van
Lierop; Chelation Prior to Hydrogen Peroxide
Bleaching of Kraft Pulps: An Overview; Journal
Pulp and Paper Science: 21, (8) J268 - J273
(1995)
[11] A. P. Johnson; Worldwide Survey of Oxygen
Bleach Plants; Non-Chlorine Bleaching Conf.
Proceedings, Hilton Head, 1992
[12] H. Loutfi; The use of hydrogen peroxide in
bleaching of kraft softwood pulp; CPPA annual
meeting 1981, B71 - B77
[13] C. A. dos Santos, H. U. Sss, O. Mambrim
Filho; Flexibilizao da sequncia de bran-
queamento ECF da Bahia Sul Celulose s. a.; 28
Congresso anual de celulose e papel, So Paulo
1995
[14] R. Reeves, R. Boman and S. Nordn; Tappi 1995
Pulping Conf. Proceedings 263 - 267
[15] N. Nimmerfroh, H. U. Sss, J. D. Kronis; Two-
stage, high brightness bleaching of sulfite pulp;
1992 Tappi Pulping Conf. Proceedings 791 - 801
[16] R. Hock, W. Czirnich; Chlorfreie Bleiche von
Zellstoff mit Einbindung von Abwasser in den
Kreislauf der Kochereiabwsser - Konzept und
Ergebnisse einer neuen Bleichsequenz; Papier
47, (10A) V24 - V29 (1993)
[17] J. Basta, L. Holtinger, L. Lundgren,and P. Fas-
ten; Tappi 1991 Intn. Pulp Bleaching Conf.
Proceedings Vol 3 , 23 - 33
[18] J. Basta, L. Holtinger, K. Gutke; Reducing Levels
of AOX, Part 4: Lignox on Hardwood Pulps; Tappi
1991 Pulping Conf. Proceedings Vol 1; 153 - 158
[19] J. Basta, L. Andersson and W. Hermansson;
The Potential of Lignox and Complementary
Combinations; Proceedings Non-Chlorine
Bleaching Conf., Hilton Head 1992
[20] H. U. Sss, N. F. Nimmerfroh and O. Mambrim
Filho; TCF bleaching of eucalyptus kraft pulp:
The selection of the right sequence and the best
conditions; Tappi 1996 Intn. Pulp Bleaching Conf.
Proceedings 253 - 260
[21] J. Basta, L. Holtinger, P. Lundgren and C.
Persson; Emerging Technologies In TCF
Bleaching; Tappi 1995 Pulping Conf. Proceed-
ings, 53 - 57
[22] L. Sjdin, S. Norden and R. Boman; Extended
Delignification with Oxygen and Hydrogen
Peroxide in ECF and TCF Sequences, Tappi
1994 Pulping Conf. Proceedings, 21 - 27
[23] J. Devenyns, R. Detroz, A. Renders; Enhanced
P stages for ECF and TCF bleaching; Tappi 1996
Intern. Pulp Bleaching Conf. Proceedings 295 -
302

Das könnte Ihnen auch gefallen